• No results found

Transposable elements contribute to fungal genes and impact fungal lifestyle

N/A
N/A
Protected

Academic year: 2021

Share "Transposable elements contribute to fungal genes and impact fungal lifestyle"

Copied!
10
0
0

Loading.... (view fulltext now)

Full text

(1)

transposable elements contribute

to fungal genes and impact fungal

lifestyle

Anna Muszewska

1

, Kamil steczkiewicz

2

, Marta stepniewska-Dziubinska

1

&

Krzysztof Ginalski

2

the last decade brought a still growing experimental evidence of mobilome impact on host’s gene

expression. We systematically analysed genomic location of transposable elements (TEs) in 625

publicly available fungal genomes from the NCBI database in order to explore their potential roles in genome evolution and correlation with species’ lifestyle. We found that non-autonomous tes and remnant copies are evenly distributed across genomes. In consequence, they also massively overlap with regions annotated as genes, which suggests a great contribution of te-derived sequences to host’s coding genome. Younger and potentially active tes cluster with one another away from genic regions. this non-randomness is a sign of either selection against insertion of tes in gene proximity or target site preference among some types of tes. proteins encoded by genes with old transposable

elements insertions have significantly less repeat and protein-protein interaction motifs but are richer

in enzymatic domains. However, genes only proximal to tes do not display any functional enrichment.

Our findings show that adaptive cases of TE insertion remain a marginal phenomenon, and the

overwhelming majority of tes are evolving neutrally. eventually, animal-related and pathogenic fungi

have more TEs inserted into genes than fungi with other lifestyles. This is the first systematic,

kingdom-wide study concerning mobile elements and their genomic neighbourhood. the obtained results should inspire further research concerning the roles tes played in evolution and how they shape the life we know today.

Transposable elements (TEs) constitute a significant but understudied fraction of eukaryotic genomes. They are mobile genetic units that proliferate and expand to distant genomic regions. TEs are classified into two classes based on their transposition mechanism. Class I groups elements that transpose using an RNA intermediate, whereas Class II members skip RNA transcript and transpose directly from DNA to DNA1. TE landscape of most

eukaryotic genomes consists of Class I representatives including: retrotransposons with Long Terminal Repeats (LTR retrotransposons), Long Interspersed Nuclear Elements (LINEs), Short Interspersed Nuclear Elements (SINEs), as well as of Class II DNA transposons that encode a classic DDE transposase (“cut and paste” DNA TEs, TIRs) or comply with yet unknown mechanism of transposition, e.g. Helitrons and Polintons/Mavericks.

For a long time, transposable elements were considered just another species of “junk DNA” and the hypothesis on their regulatory roles raised by Barbara McClintock2 remained ignored. Their impact on eukaryotic evolution

and genome function is still a matter of vigorous debate between two extremes: TEs as passive genetic material for selection on one side and powerful factors that immediately impact cell and organism’s fate on the other3,4.

Nonetheless, TEs can be considered as molecular parasites, which introduce mutations and eventually contribute significantly to genome size inflation5–7. Like other parasites, they take part in an arms race against host’s defence

mechanisms and organisms have developed multiple complex mechanisms to keep their genomes clear from for-eign DNA. The most common are DNA methylation8, targeting by tRNA-derived small RNAs9–11, RNAi mediated

silencing12 and repeat-induced point mutations13.

TE insertion breaks continuity of co-selected traits, alters gene transcription, leads to chromosomal rear-rangements by promoting recombination14,15 and promotes insertional mutations, which can impose deleterious

consequences for target loci4. In the last decade, remarkable examples of TE functional impact on host, mostly

1Institute of Biochemistry and Biophysics, Polish Academy of Sciences, Pawinskiego 5A, 02-106, Warsaw, Poland. 2Laboratory of Bioinformatics and Systems Biology, CeNT, University of Warsaw, Zwirki i Wigury 93, 02-089, Warsaw, Poland. Anna Muszewska and Kamil Steczkiewicz contributed equally. correspondence and requests for materials should be addressed to A.M. (email: musze@ibb.waw.pl)

Received: 3 October 2018 Accepted: 26 February 2019 Published: xx xx xxxx

(2)

animal, have been described, including organ development16, karyotype changes17, cell fate regulation18 and stress

response modulation19. TE-derived genes play crucial roles in all living organisms and massively alter

expres-sion of the proximal genes20. A TE can modify host transcripts via exonisation of itself, induction of original

exon skipping what leads to alternative transcripts, insertion into an ORF (into an existing frame) creating a new fusion protein, and insertion of alternative polyadenylation signals. It can also interfere with gene regu-lation by delivering novel, illegitimate promoter sequences. For example, a single transposon-derived protein, CSB-PGBD3 (domesticated transposase) can interact with as many as 900 remnant TE sequences and plays roles in gene regulation upon DNA damage21. Also, host’s protein-coding mRNAs can be occasionally retrotransposed

by retrotransposon-related machinery, which can result in formation of novel pseudogenes and genes22. The latter

might eventually donate polyadenylation sites to neighbouring genes and further expand transcript diversity22.

Phenomena resulting from gene-transposable element proximity have been thoroughly studied mainly for model animals20,23,24 and plants3, and only a few studies included fungal genomes despite genomic resource

abun-dance25–27. For instance, a remnant LTR retrotransposon insertion into promoter region of a gene coding for

MFS1 transporter was found to induce this gene overexpression and to enhance fungicide resistance28. Also, gene

clusters can be regulated by neighbouring TEs, e.g. the penicillin cluster in Aspergillus nidulans has lower expres-sion in the absence of Pbla element29. In Schizosaccharomyces pombe, Tf1 element has a preference for promoters

of stress-related genes, which eventually enhances their expression and promotes survival of the fungus30.

TE neighbourhood within a window of 1 kb has a repressive effect on neighbouring genes in fungi equipped with functional methylation machinery, but casts no such effect in Saccharomyces cerevisiae, which lacks methyl-ation25. Genes within 1 kb to a Gypsy or hAT transposon have lower expression in Coccidioides immitis31. In this

organism, TEs are often inserted in proximity of phosphorylation-related genes. Castanera and colleagues showed also that the presence of TE clusters has more pronounced regulatory effects on gene expression as compared to a single TE upstream or downstream25.

Some fungal pathogens of plants have genomes with a clearly dualistic architecture described by the two-speed model of evolution. The core genome is densely packed with housekeeping genes while a lifestyle-adapting part contains effector genes and TEs7,32. This genome architecture was reported for versatile fungal pathogens among

them Fusarium33, Leptosphaeria34 and Verticillium35. The lifestyle-specific genome is expected to be enriched in

TEs, as they may play roles in host switching and adaptation to new ecological niches36, which can be observed in

Magnaporthe oryzae, where genes associated with TEs are involved in host specialization37. In consequence, even

closely related fungal taxa may differ significantly in transposable content, e.g. Amanita species with saprophytic and mycorrhizal lifestyles38.

Encouraged by the aforementioned experimental screenings demonstrating the impact of TEs on gene expres-sion, we performed a systematic analysis of their genomic context in publicly available fungal genomes. Here, we investigate the immediate neighbourhood of transposable elements, with special focus on co-localizing genes. Moreover, we interpret our results from a lifestyle perspective.

Methods

Genomes and transposable elements.

Fungal proteomes were downloaded from NCBI on 17th August 201639 and genomic sequences were downloaded from NCBI genome portal on 18th of August 2016. 625 genomic

assemblies with corresponding proteomes analysed in this study are listed in Supplementary Table S1. Genome sequences deposited at the NCBI were obtained using diverse sequencing techniques, with different sequencing depths, assembled and annotated using a plethora of approaches. In consequence, there ought to be gene calling inconsistencies, missing genome fragments and to deal with it our study will focus on general trends instead of singularities. Genomic coordinates of TEs were inferred in the course of de novo and homology-based TE anno-tation. Irf inverted repeat finder40 (irf parameters used: matching weight 2, mismatching weight 3, indel penalty

5, match probability 80, indel probability 10, minimum alignment score to report 20, maximum stem length to report 500000, MaxLoop 10000, additional options: -a3 -t4 1000 -t5 5000) and RepeatModeler41 were used to

detect TE candidates de novo. Irf hits were classified using the RepeatModeler annotating script. Multiple overlap-ping hits were removed by clustering with RepBase database entries42 using CD-HIT43, and the resulting sequence

dataset of TE consensus sequences was used as a library for RepeatMasker homology search44 (RepeatMasker was

invoked with options: -gccalc -no_is, TEs with scores above 200 were taken). All the resulting sequences were scanned with manually curated list of reference Pfam HMM profiles (using pfam_scan.pl with E-value threshold 0.01)45 and CDD profiles (RPS-BLAST with E-value threshold 0.001)46 listed in Supplementary Table S2. This TE

annotation pipeline has been successfully employed previously in the study of DNA TE’s47 as well as in a growing

number of genome annotation studies48–50. The chosen protein domains are either associated with TE activity or

related to TEs and were collected based on TE architectures known from RepBase and literature. The elements containing sequences similar to known TE-related domains are labelled along the manuscript as “with domain” transposable elements. Sequences without detectable similarity to known TE domains were considered as frag-ments and remnants of old TEs. A schematic workflow of the analyses is shown as Supplementary Fig. S1.

Neighbourhood classification.

Three classes of TE neighbours were defined: (i) nothing, (ii) other TE and (iii) gene. In order to provide a robust and consistent neighbourhood classification, we defined the following set of rules. First of all, to adhere to varying genome architectures for each species, an adaptive scanning window size was estimated as a median of gene distances in the whole assembly (Supplementary Table S1) with the top size of 1 kb. The minimal median gene distance value was 71 for Enterocytozoon bieneusi and maximum 8,997 for Edhazardia aedis. In total, 12 analysed assemblies had the window narrower than 100 bp while 79 - wider than 1 kb. All protein sequences encoded by genes partially overlapping with TE coordinates were scanned against a list of TE-related protein domains using pfam_scan.pl tool. If the gene had no detectable TE-related domains, the TE borders were shortened and the gene became TE’s immediate neighbour; otherwise, the gene was included into

(3)

TE’s borders and the neighbourhood was determined against expanded TE coordinates. If TE fully covered any gene, it replaced this gene in further neighbourhood assessment. Moreover, if the neighbouring gene contained an inner TE, which was also located within the window distance to the analysed TE, this inner element was anno-tated as a neighbour (Supplementary Fig. S2). When two or more TEs overlapped, they were merged together and tagged with the most specific annotation common to all of participating TEs. When merged TEs were of totally distinct species, the newly defined TE was tagged as a ‘composite’.

A TE inserted into a gene can reside within a 3′ UTR, 5′ UTR, intron or exon. Unfortunately, the majority of analysed assemblies lacked gene inner structures and even less included UTRs at all. In consequence, we were not able to study the detailed location of TEs at a sub-genic level.

The encoded proteins were scanned for secretion signals using TargetP51 and were assigned to GO categories

using pfam2GO table52.

Data analysis.

All genomes with incomplete annotation, for instance without gene predictions, were excluded from analysis, as mentioned above. Genome statistics (size, density, intron per gene) were computed based on the assembly sequences and gff annotation files downloaded from the NCBI database. Since gene call-ing strategies vary in reliability between genomes and initial data quality directly impacts our neighbourhood analyses, we have selected only highly significant patterns emerging from analyses described in this manuscript.

Information on fungal lifestyles, as in our previous study on DNA transposons47, was derived from the

availa-ble literature. Categories including host type (plant, animal, fungus), main habitat (soil/dung, water) and lifestyle (pathogenic, symbiotic and saprotrophic) were assigned to every species in the dataset. Noteworthy, a single fun-gus could represent multiple categories, if applicable, e.g. species functioning both as a plant symbiont and animal pathogen (see Supplementary Table 1). Taxonomical annotation was derived from the NCBI taxonomy database, with manual curation when needed (see Supplementary Table 1). TE types were described using a 2-level hierar-chy comprising Wicker’s orders/Repbase classes (e.g. LINEs, SINEs, LTRs) and superfamilies (e.g. Copia, hAT).

Exploratory analysis and basic statistics for the dataset were carried out using pandas and seaborn Python packages. Statistical tests were performed in Python with the scipy package. Distributions of distances between TEs and genes for fungi with different lifestyles were compared with Mann-Whitney U test. Relationships between the number of TE inserted into genes as well as other genome statistics were evaluated using McFadden’s R-squared for logistic regression with binomial errors. The logistic regression models were built with statsmodels package. Enrichment analyses were performed using binomial distribution, and upper-bounds for p-values were computed with formula derived from Hoeffding’s inequality:

≤    − −     ‐ p value exp np k n 2( ) ,2

where n is the number of trials, k is the number of successes and p is the success probability.

The genome features are available as Supplementary Table 1 and the code for statistical procedures is available as a python code in a Jupyter Notebook (Supplementary File 1)53.

Results

te fragments contribute to fungal genes.

2,023,812 TE fragments and 293,746 TEs with a TE-related protein domain were identified in 625 fungal genomes (Supplementary Table S2 lists protein domains either asso-ciated with TE activity or related to TEs). Our TE counts are likely to be underestimated and there are two major reasons for that. The first and more fundamental one is a derivative of methods used in whole genome sequencing projects, which rely mostly on sequencing reads of lengths insufficient for effective reconstitution of long repeat regions. The second reason lies in our approach, as we chose to apply rather stringent filtering of identified TE fragments in order to increase the method’s reliability. All TE candidates had to be confirmed with RepeatMasker using extended fragments library as described in Methods section. Additionally, all TEs regarded as still func-tional were supposed to contain at least one known TE-related protein domain.

Fungal genomes have different gene densities and architectures ranging from very compact in endoparasitic Microsporidia to relatively big and complex Tuber and Puccinia genomes. A question arises whether and how such rough genome characteristics correlate with TE localization in different taxa.

We found that non-autonomous TEs and remnants massively overlap with regions annotated as genes. These results suggest a great contribution of TE-derived sequences to host’s genes (Fig. 1). 50.6% of non-autonomous TEs are inserted into a genic region (1,024,918) and 11.6% (235,593) TEs fragments were found in proximity of gene on either side, being equally ubiquitous downstream and upstream of genes (116,722 downstream, 118,871 upstream). The location of a TE fragment between two genes is relatively rare (1.8% of TEs, 36,841). That totals to 64% of non-autonomous TEs co-localising with genes and points at the compact architecture of many fungal genomes assuming random distribution of ancient TEs and genes in many of the fungal genomes. More compact genomes host more remnant TEs inserted into genes as compared to genomes with greater non-genic space (Fig. 2A). 14.6% of TEs had another TE as a neighbour either upstream (147,874) or downstream (147,114), 11.8% (238,024) TEs were located in-between other TEs, while 9.6% of TEs fragments (193,448) had neither genes nor TEs identified within the chosen scanning window. In total, 35.9% of analysed TEs either had another TEs as exclusive neighbours or lacked neighbourhood at all.

Active tes cluster with other tes.

Transposable elements with at least one protein domain typical for mobile elements have a distinct genomic neighbourhood profile. They are rarely found within or in close prox-imity of genes (less than 15.9%, 46,789 of these elements are inserted into a gene and 16,2%, 47,493 are close to a gene) and tend to cluster with other TEs (almost 49.5%, 145,384) or locate in regions without genes and other TEs

(4)

(18.4%, 54,080). Academ is an exception here because most of their copies contain recognizable protein domains themselves (81%, 1,568/1,938) and are classified as overlapping with a gene. These domains are encoded by a TE but are not classified as TE-specific (e.g. DEAD/DEAH box helicase domain (PFAM: PF00270) Replication, recombination and repair, recQ_fam (CDD:129701)), which hacks classification criteria and eventually makes Academs frequently annotated as host genes.

This non-random distribution of potentially active TEs might be a sign of general negative selection imposed on TEs interfering with gene coding regions or target site preference as observed for some types of TEs (e.g. Zisupton54). Even if insertion preference might play a pivotal role in shaping the genomic landscape of active

elements, once they became inactivated, the evolutionary pressure against them faded and TE fragments have survived in genomic areas where active TEs are not allowed.

Genome properties and te localization.

There is a significant correlation (R2

McF= 0.53 for TEs with

a domain and R2

McF= 0.65 for TE fragments) between the fraction of TEs targeting genes and genome

com-pactness measured as fraction of the genome occupied by genes (Fig. 2A,B). The smaller the gene distances and fewer non-genic regions, the more TE-related sequences overlap with genes likely as a result of scarcity of other genomic locations.

Overall ubiquity of remnant TEs in gene neighbourhood can be a consequence of the random distribution of TEs resulting from neutrality of old and fragmented TEs, lack of traceable target site preference among most types of TEs, and most probably recurrent usage of ancient TE-derived sequences. Interestingly, we observe a binomial distribution of in-gene insertion frequency for TEs with TE-related domains (Fig. 2C). The two peaks correspond to two distinct genome architectures within fungi: one with a higher fraction of both remnant and coding TEs in genes (mostly in Saccharomycotina, see Supplementary Fig. S3) and the other one with only remnant TE debris

(5)

located within genes (filamentous fungi). The former TE distribution is peculiar and might be a consequence of Saccharomycotina’s selection on compactness of the genomes.

Remnant tes populate enzyme-encoding genes.

Non-autonomous TEs and TE rem-nants. Protein-coding genes impacted by old TE insertions are significantly depleted in protein repeat motifs such as Ankyrin, WD40 and in protein-protein interaction domains like F-box (See Supplementary Table 3). This pattern has not been described so far and will be explored in detail in further studies. One might expect that

Figure 2. Distribution of in-gene TEs with (A) and without (B) TE-related domains in relation to genomes

with different gene density. (C) Distribution of genome assemblies with different fractions of TEs located in genes.

(6)

repeat sequences will appear as artefacts with de novo TE searches mainly due to large families present in a single genome. However, the obtained result showing protein repeat underrepresentation can be a hallmark of method robustness and supports the lack of such artefacts, at least manifested at protein-level. Additionally, it might suggest previously undescribed selection pattern yet to be understood. Fragments of LINEs co-localise with ATP-synt_ab_N ATP synthases (PF02874) and Metallophos phosphoesterases (PF00149). Non-autonomous LTR retrotransposons are preferentially associated with genes coding for Aconitase (PF00330), Catalase (PF00199), Peptidase_M41 (PF01434) and Chitin_synth_1 synthase (PF01644). Remnants of DNA TE are found with genes coding for Glyco_hydro_3_C hydrolase (PF01915) and PNP_UDP_1 phosphorylases (PF01048). Helitron remainings can be found in proteins with Peptidase_S8 (PF00082) and Pkinase (PF00069) domains.

TEs with a coding region. Functional transposable elements rarely insert into genes and do not show a statis-tically significant preference for specific protein domains. Usually, they cluster with other TEs in genomic areas containing fewer genes. Eventually, genes infested by them often carry TE-related domains, and are likely to be TEs misannotated as genes and included into proteomes. TEs tend to insert into other TEs leading to the forma-tion of TE-clusters or composite elements55,56.

TE location, abundance and hosts’ ecology. Animal-related and pathogenic fungi have more TEs inserted into genes as compared to fungi with other lifestyles (Fig. 3). Plant-related, saprotrophic organisms and those living in soil or on dung have fewer TEs overlapping with genic regions. This effect is straightforwardly correlated with genome compactness of animal pathogenic fungi and genome expansion present in many plant-associated fungi7. Genome architecture seems to be the dominant factor determining the relationship between the coding

and non-coding genome. Plant-associated fungi have a greater average distance between TEs and genes (370 bp) and fewer genes close to TEs as compared to non-plant related fungi (351 bp between gene and TE on average, p-value = 7.7e-78). Both features are likely attributed to greater genome sizes and overall decrease in gene density. Small secreted proteins. Small secreted proteins (SSPs) are often related with plant-associated lifestyle providing effector activity modulating host performance57. Plant-pathogenic fungi are known for their peculiar genome

architecture with fast-evolving genomic regions rich in repeat proteins, SSPs and TEs7. We tested whether SSPs

would indeed cluster with TEs in terms of the neighbourhood defined in this paper. The SSPs were defined either as shorter than 300 amino acids and predicted to be secreted or additionally possessing more than 5% of cysteines (which narrowed the gene list). Regardless of the applied definition, we found no statistical support for the associ-ation between SSP neighbourhood with TEs. One of the possible explanassoci-ations here would be that in fast-evolving genome parts, the maximum distance allowing for TE-gene influence might be bigger than 1 kb averaged for many genomes, with a majority of them having a more uniform genome architecture. Our analyses are also lim-ited by assembly fragmentation particularly affecting repeat-rich genome regions. SSPs understood as genes cod-ing short secreted proteins constituted 3.6% of all neighbourcod-ing genes and 6.1% of all protein codcod-ing genes. These values varied among genomes with Agaricomycetes (n = 72, mean of 78) having more SSPs in TE neighbourhood than Tremellomycetes (n = 31, mean of 10). Among Pezizomycotina, Eurotiomycetes had less SSPs co-localising with TEs (n = 122, mean 39) than Dothideomycetes (n = 39, mean 69) and Leotiomycetes (n = 33, mean 103), being the most SSP rich in proximity of TEs.

Discussion

The aim of this study was to explore neighbourhood of fungal transposable elements, either functional or not. TEs are intrinsically linked to genome evolution and constitute minor but still ubiquitous fraction of most fungal genomes. Their roles as potent regulatory elements, genomic parasites and nearly neutral sequences are being revised constantly58. According to Arkhipova and others, the most transposable elements remain silent, evolve

in a neutral fashion and only a minor fraction gets ever involved in adaptive roles59. Our results seem to confirm

this perspective, showing no correlation between TE neighbourhood and gene function for many TE families and remnant elements. We might not be able to detect rare events on this large scale e.g. a new regulatory network that uses TEs as TF binding sites. With the advancement of single cell sequencing technologies, it will soon become feasible to observe TE movements and distribution across fungal populations without being limited to model organisms only.

Observed localisation of TEs in 625 fungal genomes shows a dichotomy between relatively young elements depleted in genes, and remnant sequences clearly derived from transposable elements, now more deteriorated, which are equally likely to be found both, within genes and in other locations. This phenomenon provides a path-way to exaptation of TEs, producing new coding regions and utilization as evolutionary raw material for selection. The significance of exaptation in the course of Metazoa evolution has been noted by Scharder and Schmitz in their review on TEs in adaptive evolution58.

There are numerous factors shaping TEs distribution ranging from target site preferences in some retrotrans-posons favouring insertion upstream of polymerase III transcribed genes60, strand preference in LTR

retrotrans-posons, via genome rearrangements, to forces of selection and genetic drift acting at a population scale removing TEs with deleterious phenotypes56,61. Regardless of insertional preference, present in some TE types, the overall

pattern of genomic distribution of both functional and dead elements corroborates a random fashion of TE dis-persal within genomes. These genomic parasites remain active outside of genes, where they are less likely to cause deleterious mutations. TEs with a coding region are predominantly in a distance from host genes, what might be related to the repressive effect of many TEs on neighbouring genes31. Remnant TEs are not subjected to such

constraints and can now be used as raw material for new coding sequences.

The proportion of the genome originated from TEs varies in different fungal lineages as shown previously7,47.

(7)

pattern suggests the presence of constraints imposed on the size of small genomes, despite multiplication of TEs and randomness of the insertion process. In consequence, small genomes remain small, and large ones grow. The growth of big fungal genomes can be acknowledged to genetic drift, they change with time gaining new slightly deleterious mutations, mobile elements and introns62–64. In contrary, the very compact genomes of yeast-like

organisms are likely a result of selection62,65.

Genome architecture seems to depend on fungus ecology. Most fungi with complex genomes shaped by numerous TEs are plant-associated which has been noticed previously7,47. Plant-related fungi are known to

use SSPs to deal with plant’s immune reaction. It has been claimed that SSP-coding genes co-localise with TEs,

Figure 3. Distribution of TEs in fungi with a given lifestyle. Significance of differences is assessed with Mann–

(8)

however, we did not observe this effect. The latter effect can be masked by the underrepresentation and frag-mentation of repeat rich genomic locations in assemblies. Surprisingly, our findings point at several previously unreported correlations between occurrence of TE-gene overlapping and animal-related and/or pathogenic host lifestyle. It remains an open question whether there is a causative relation between fungal ecology and TEs dis-tribution in the genome – it may be validated by experiments involving multiple high-quality genomes and tran-scriptomes from closely related taxa differing in lifestyle.

Analysis of an extensive dataset of genomes covering organisms of diverse genome sizes, lifestyles, taxonomic position and TE abundance enabled us to ask whether TE insertions are linked to specific functional categories described previously, such as stress response66, mutualism67 or phosphorylation31. We found no general

rela-tionship between the aforementioned biological functions and TE neighbourhood. This finding may suggest that these phenomena are taxon-specific. However, we did find associations between TEs and several unrelated enzyme classes, for particular fungal lineages and TEs classes. Our conclusion supports Arkhipova’s hypothesis that adaptive roles of TEs will remain statistically undetectable and will remain a case-by-case phenomenon. We might hypothesise that TEs can play diverse roles, including adaptive ones, in the course of evolution of particular fungal populations, each being shaped by its constraints. When analysed together, these specific cases are masked by the dominant random and neutral fashion of TE evolution.

Data Availability

Information processed in statistical analyses is available as Python code and Excel tables.

References

1. Wicker, T. et al. A unified classification system for eukaryotic transposable elements. Nat. Rev. Genet. 8, 973–982, https://doi. org/10.1038/nrg2165 (2007).

2. Mc, C. B. The origin and behavior of mutable loci in maize. Proc Natl Acad Sci USA 36, 344–355 (1950).

3. Bennetzen, J. L. & Wang, H. The contributions of transposable elements to the structure, function, and evolution of plant genomes. Annu. Rev. Plant Biol. 65, 505–530, https://doi.org/10.1146/annurev-arplant-050213-035811 (2014).

4. Werren, J. H. Selfish genetic elements, genetic conflict, and evolutionary innovation. Proceedings of the National Academy of Sciences of the United States of America 108(Suppl 2), 10863–10870, https://doi.org/10.1073/pnas.1102343108 (2011).

5. Rodriguez, F. & Arkhipova, I. R. Transposable elements and polyploid evolution in animals. Curr Opin Genet Dev 49, 115–123,

https://doi.org/10.1016/j.gde.2018.04.003 (2018).

6. Wendel, J. F., Lisch, D., Hu, G. & Mason, A. S. The long and short of doubling down: polyploidy, epigenetics, and the temporal dynamics of genome fractionation. Curr Opin Genet Dev 49, 1–7, https://doi.org/10.1016/j.gde.2018.01.004 (2018).

7. Moller, M. & Stukenbrock, E. H. Evolution and genome architecture in fungal plant pathogens. Nat Rev Microbiol 15, 756–771,

https://doi.org/10.1038/nrmicro.2017.76 (2017).

8. Zemach, A., McDaniel, I. E., Silva, P. & Zilberman, D. Genome-Wide Evolutionary Analysis of Eukaryotic DNA Methylation. Science 328, 916–919, https://doi.org/10.1126/science.1186366 (2010).

9. Wang, Q. et al. The tRNA-Derived Small RNAs Regulate Gene Expression through Triggering Sequence-Specific Degradation of Target Transcripts in the Oomycete Pathogen Phytophthora sojae. Front. Plant Sci. 07, https://doi.org/10.3389/fpls.2016.01938

(2016).

10. Martinez, G., Choudury, S. G. & Slotkin, R. K. tRNA-derived small RNAs target transposable element transcripts. Nucleic Acids Res. 45, 5142–5152, https://doi.org/10.1093/nar/gkx103 (2017).

11. Schorn, A. J., Gutbrod, M. J., LeBlanc, C. & Martienssen, R. LTR-Retrotransposon Control by tRNA-Derived Small RNAs. Cell 170, 61–71.e11, https://doi.org/10.1016/j.cell.2017.06.013 (2017).

12. Torres-Martínez, S. & Ruiz-Vázquez, R. M. The RNAi Universe in Fungi: A Varied Landscape of Small RNAs and Biological Functions. Annu. Rev. Microbiol., https://doi.org/10.1146/annurev-micro-090816-093352 (2017).

13. John Clutterbuck, A. Genomic evidence of repeat-induced point mutation (RIP) in filamentous ascomycetes. Fungal Genet. Biol. 48, 306–326, https://doi.org/10.1016/j.fgb.2010.09.002 (2011).

14. Laricchia, K. M., Zdraljevic, S., Cook, D. E. & Andersen, E. C. Natural Variation in the Distribution and Abundance of Transposable Elements Across the Caenorhabditis elegans Species. Mol Biol Evol 34, 2187–2202, https://doi.org/10.1093/molbev/msx155 (2017). 15. Kent, T. V., Uzunovic, J. & Wright, S. I. Coevolution between transposable elements and recombination. Philos Trans R Soc Lond B

Biol Sci 372, https://doi.org/10.1098/rstb.2016.0458 (2017).

16. Erwin, J. A., Marchetto, M. C. & Gage, F. H. Mobile DNA elements in the generation of diversity and complexity in the brain. Nat. Rev. Neurosci. 15, 497–506, https://doi.org/10.1038/nrn3730 (2014).

17. Carbone, L. et al. Gibbon genome and the fast karyotype evolution of small apes. Nature 513, 195–201, https://doi.org/10.1038/ nature13679 (2014).

18. Zemojtel, T., Kielbasa, S. M., Arndt, P. F., Chung, H.-R. & Vingron, M. Methylation and deamination of CpGs generate p53-binding sites on a genomic scale. Trends Genet. 25, 63–66, https://doi.org/10.1016/j.tig.2008.11.005 (2009).

19. Rey, O., Danchin, E., Mirouze, M., Loot, C. & Blanchet, S. Adaptation to Global Change: A Transposable Element–Epigenetics Perspective. Trends Ecol. Evol. 31, 514–526, https://doi.org/10.1016/j.tree.2016.03.013 (2016).

20. Cowley, M. & Oakey, R. J. Transposable elements re-wire and fine-tune the transcriptome. PLoS Genet. 9, e1003234, https://doi. org/10.1371/journal.pgen.1003234 (2013).

21. Bailey, A. D. et al. The conserved Cockayne syndrome B-piggyBac fusion protein (CSB-PGBD3) affects DNA repair and induces both interferon-like and innate antiviral responses in CSB-null cells. DNA Repair 11, 488–501, https://doi.org/10.1016/j. dnarep.2012.02.004 (2012).

22. Vinckenbosch, N., Dupanloup, I. & Kaessmann, H. Evolutionary fate of retroposed gene copies in the human genome. Proc. Natl. Acad. Sci. USA 103, 3220–3225, https://doi.org/10.1073/pnas.0511307103 (2006).

23. del Rosario, R. C. H., Rayan, N. A. & Prabhakar, S. Noncoding origins of anthropoid traits and a new null model of transposon functionalization. Genome Res. 24, 1469–1484, https://doi.org/10.1101/gr.168963.113 (2014).

24. Rebollo, R., Romanish, M. T. & Mager, D. L. Transposable elements: an abundant and natural source of regulatory sequences for host genes. Annu. Rev. Genet. 46, 21–42, https://doi.org/10.1146/annurev-genet-110711-155621 (2012).

25. Castanera, R. et al. Transposable Elements versus the Fungal Genome: Impact on Whole-Genome Architecture and Transcriptional Profiles. PLoS Genet. 12, e1006108, https://doi.org/10.1371/journal.pgen.1006108 (2016).

26. Santana, M. F. et al. Abundance, distribution and potential impact of transposable elements in the genome of Mycosphaerella fijiensis. BMC Genomics 13, 720, https://doi.org/10.1186/1471-2164-13-720 (2012).

27. Santana, M. F. et al. Characterization and potential evolutionary impact of transposable elements in the genome of Cochliobolus heterostrophus. BMC Genomics 15, 536, https://doi.org/10.1186/1471-2164-15-536 (2014).

(9)

28. Omrane, S. et al. Plasticity of the MFS1 Promoter Leads to Multidrug Resistance in the Wheat Pathogen Zymoseptoria tritici. mSphere 2, https://doi.org/10.1128/mSphere.00393-17 (2017).

29. Shaaban, M. et al. Involvement of transposon-like elements in penicillin gene cluster regulation. Fungal Genet. Biol. 47, 423–432,

https://doi.org/10.1016/j.fgb.2010.02.006 (2010).

30. Feng, G., Leem, Y.-E. & Levin, H. L. Transposon integration enhances expression of stress response genes. Nucleic Acids Res. 41, 775–789, https://doi.org/10.1093/nar/gks1185 (2013).

31. Kirkland, T., Muszewska, A. & Stajich, J. Analysis of Transposable Elements in Coccidioides Species. Journal of Fungi 4, 13, https:// doi.org/10.3390/jof4010013 (2018).

32. Dong, S., Raffaele, S. & Kamoun, S. The two-speed genomes of filamentous pathogens: waltz with plants. Curr. Opin. Genet. Dev. 35, 57–65, https://doi.org/10.1016/j.gde.2015.09.001 (2015).

33. Sperschneider, J. et al. Genome-Wide Analysis in Three Fusarium Pathogens Identifies Rapidly Evolving Chromosomes and Genes Associated with Pathogenicity. Genome Biol. Evol. 7, 1613–1627, https://doi.org/10.1093/gbe/evv092 (2015).

34. Rouxel, T. & Balesdent, M.-H. Life, death and rebirth of avirulence effectors in a fungal pathogen of Brassica crops, Leptosphaeria maculans. New Phytol. 214, 526–532, https://doi.org/10.1111/nph.14411 (2017).

35. Faino, L. et al. Transposons passively and actively contribute to evolution of the two-speed genome of a fungal pathogen. Genome Res. 26, 1091–1100, https://doi.org/10.1101/gr.204974.116 (2016).

36. Ma, L.-J. et al. Comparative genomics reveals mobile pathogenicity chromosomes in Fusarium. Nature 464, 367–373, https://doi. org/10.1038/nature08850 (2010).

37. Yoshida, K. et al. Host specialization of the blast fungus Magnaporthe oryzae is associated with dynamic gain and loss of genes linked to transposable elements. BMC Genomics 17, https://doi.org/10.1186/s12864-016-2690-6 (2016).

38. Szczepaniska, M., Muszewska, E., Szprynger, K. & Niwinska-Faryna, B. Systemic lupus erythematosus and pregnancy. Wiad Lek 61, 161–165 (2008).

39. Database Resources of the National Center for Biotechnology Information. Nucleic Acids Res 45, D12–D17, https://doi.org/10.1093/ nar/gkw1071 (2017).

40. Warburton, P. E., Giordano, J., Cheung, F., Gelfand, Y. & Benson, G. Inverted repeat structure of the human genome: the X-chromosome contains a preponderance of large, highly homologous inverted repeats that contain testes genes. Genome Res. 14, 1861–1869, https://doi.org/10.1101/gr.2542904 (2004).

41. Hubley, R. & Smit, A. RepeatModeler Open-1.0, http://www.repeatmasker.org (2015).

42. Kapitonov, V. V. & Jurka, J. A universal classification of eukaryotic transposable elements implemented in Repbase. Nat. Rev. Genet. 9, 411–412; author reply 414, https://doi.org/10.1038/nrg2165-c1 (2008).

43. Fu, L., Niu, B., Zhu, Z., Wu, S. & Li, W. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinformatics 28, 3150–3152, https://doi.org/10.1093/bioinformatics/bts565 (2012).

44. Smit, A., Hubley, R. & Green, P. RepeatMasker Open-4.0, http://www.repeatmasker.org (2015).

45. Finn, R. D. et al. The Pfam protein families database: towards a more sustainable future. Nucleic Acids Res. 44, D279–285, https://doi. org/10.1093/nar/gkv1344 (2016).

46. Ncbi Resource Coordinators. Database Resources of the National Center for Biotechnology Information. Nucleic Acids Res. 45, D12–D17, https://doi.org/10.1093/nar/gkw1071 (2017).

47. Muszewska, A., Steczkiewicz, K., Stepniewska-Dziubinska, M. & Ginalski, K. Cut-and-paste transposons in fungi with diverse lifestyles. Genome Biol. Evol, https://doi.org/10.1093/gbe/evx261 (2017).

48. Gryganskyi, A. P. et al. Phylogenetic and Phylogenomic Definition of Rhizopus Species. G3 (Bethesda) 8, 2007–2018, https://doi. org/10.1534/g3.118.200235 (2018).

49. Schoville, S. D. et al. A model species for agricultural pest genomics: the genome of the Colorado potato beetle, Leptinotarsa decemlineata (Coleoptera: Chrysomelidae). Sci Rep 8, 1931, https://doi.org/10.1038/s41598-018-20154-1 (2018).

50. Teixeira, M. M. et al. Exploring the genomic diversity of black yeasts and relatives (Chaetothyriales, Ascomycota). Stud Mycol 86, 1–28, https://doi.org/10.1016/j.simyco.2017.01.001 (2017).

51. Emanuelsson, O., Nielsen, H., Brunak, S. & von Heijne, G. Predicting subcellular localization of proteins based on their N-terminal amino acid sequence. J. Mol. Biol. 300, 1005–1016, https://doi.org/10.1006/jmbi.2000.3903 (2000).

52. Mitchell, A. et al. The InterPro protein families database: the classification resource after 15 years. Nucleic Acids Res. 43, D213–221,

https://doi.org/10.1093/nar/gku1243 (2015).

53. Kluyver, T. et al. Jupyter Notebooks—a publishing format for reproducible computational workflows. Positioning and Power in Academic Publishing: Players, Agents and Agendas, 87 (2016).

54. Iyer, L. M. et al. Lineage-specific expansions of TET/JBP genes and a new class of DNA transposons shape fungal genomic and epigenetic landscapes. Proceedings of the National Academy of Sciences 111, 1676–1683, https://doi.org/10.1073/pnas.1321818111

(2014).

55. Fedoroff, N. V. Presidential address. Transposable elements, epigenetics, and genome evolution. Science 338, 758–767 (2012). 56. Levin, H. L. & Moran, J. V. Dynamic interactions between transposable elements and their hosts. Nat Rev Genet 12, 615–627, https://

doi.org/10.1038/nrg3030 (2011).

57. Kim, K.-T. et al. Kingdom-Wide Analysis of Fungal Small Secreted Proteins (SSPs) Reveals their Potential Role in Host Association. Front. Plant Sci. 7, 186, https://doi.org/10.3389/fpls.2016.00186 (2016).

58. Schrader, L. & Schmitz, J. The impact of transposable elements in adaptive evolution. Mol. Ecol, https://doi.org/10.1111/mec.14794

(2018).

59. Arkhipova, I. R. Neutral Theory, Transposable Elements, and Eukaryotic Genome Evolution. Mol. Biol. Evol. 35, 1332–1337, https:// doi.org/10.1093/molbev/msy083 (2018).

60. Sultana, T., Zamborlini, A., Cristofari, G. & Lesage, P. Integration site selection by retroviruses and transposable elements in eukaryotes. Nat Rev Genet 18, 292–308, https://doi.org/10.1038/nrg.2017.7 (2017).

61. Bourque, G. et al. Ten things you should know about transposable elements. Genome Biol 19, 199, https://doi.org/10.1186/s13059-018-1577-z (2018).

62. Kelkar, Y. D. & Ochman, H. Causes and consequences of genome expansion in fungi. Genome Biol Evol 4, 13–23, https://doi. org/10.1093/gbe/evr124 (2012).

63. Lynch, M. & Conery, J. S. The origins of genome complexity. Science 302, 1401–1404, https://doi.org/10.1126/science.1089370

(2003).

64. Lynch, M., Bobay, L. M., Catania, F., Gout, J. F. & Rho, M. The repatterning of eukaryotic genomes by random genetic drift. Annu Rev Genomics Hum Genet 12, 347–366, https://doi.org/10.1146/annurev-genom-082410-101412 (2011).

65. Todd, R. T., Forche, A. & Selmecki, A. Ploidy Variation in Fungi: Polyploidy, Aneuploidy, and Genome Evolution. Microbiol Spectr 5, https://doi.org/10.1128/microbiolspec.FUNK-0051-2016 (2017).

66. Krishnan, P. et al. Transposable element insertions shape gene regulation and melanin production in a fungal pathogen of wheat. BMC Biol. 16, https://doi.org/10.1186/s12915-018-0543-2 (2018).

67. Hess, J. et al. Transposable element dynamics among asymbiotic and ectomycorrhizal Amanita fungi. Genome Biol Evol 6, 1564–1578, https://doi.org/10.1093/gbe/evu121 (2014).

(10)

Acknowledgements

This work was supported by the National Science Centre (2017/25/B/NZ2/01880 to AM and 2014/15/B/ NZ1/03357 to KG). AM is supported by a L’Oreal Poland – UNESCO Scholarship for Women in Science. KG was also supported by the Foundation for Polish Science (TEAM). We thank Marcin Grynberg for inspiring discussions.

Author Contributions

A.M. designed the study, A.M. and K.S. prepared the data set, developed software and performed genome analyses, M.S.-D. performed statistical analyses, and A.M., K.S., M.S.-D. and K.G. interpreted the data and wrote the manuscript.

Additional Information

Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-019-40965-0.

Competing Interests: The authors declare no competing interests.

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and

institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International

License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-ative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not per-mitted by statutory regulation or exceeds the perper-mitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

References

Related documents

Key words: iatrogenic vessel injury, arteriovenous fistula, pseudoaneurysm of the common iliac artery, endovascular treatment.. Endovascular treatment of iatrogenic

The main difference between Romer‟s methodology and the sensor association rule proposed in [5], is in type of events in the discovered pattern, in which he targeted all

All pregnant women with GDM underwent clinical evaluation and answered a questionnaire with socio- demographic information such as age, self-reported eth- nicity (White or

Wu and Tsai (2007), in a study they conducted to interpret the information searched on the Web and information searching strategies, collected data from 1220 students via the

In conclusion, in a nationally representative sample of individuals with albuminuria and preserved eGFR, both uninsurance and public insurance were associated with increased

In the "Theory of Modes" Frye says that "In low mimetic tragedy, pity and fear are neither purged nor absorbed into pleasures, but are communicated externally,

El Mir, Seboui (2006) in their study on corporate governance and earnings management and the relationship between economic value added and shareholders’ created wealth concluded

Self-rated health and smoking among physicians and general population with higher education in Estonia: results from cross-sectional studies in 2002 and 2014.. Kersti Pärna * ,